Angular Momentum Algebra. Abstract

Size: px
Start display at page:

Download "Angular Momentum Algebra. Abstract"

Transcription

1 Angular Momentum Algebra Michael W. Kirson Department of Particle Physics Weizmann Institute of Science Rehovot, Israel (Dated: January 2002) Abstract This brief summary of the quantum theory of angular momentum is intended as a heuristic, reasonably self-contained presentation of useful results. It has no pretensions to rigour and certainly does not presume to supplant the existing detailed texts on the subject, which should be consulted for a more thorough treatment of the topics touched on below. A brief bibliography is given at the end. In addition, most texts on quantum mechanics have at least some dicussion of the quantum mechanics of angular momentum, some of them quite extensive. 1

2 I. BASIC DEFINITIONS In classical mechanics, the angular momentum of a point object is defined as the vector product of its position and momentum vectors, L = r p. In quantum mechanics, where r and p are operators, one for each component of each vector, this same definition produces a set of three operators, L x, L y and L z. From the standard commutation relation [r α, p β ] = i hδ αβ, where the subscripts α and β denote the indices x, y or z and δ αβ is the Krönecker delta, it follows that [L x, L y ] = i hl z (1.1) and similarly for any cyclic permutation of the indices x, y and z. These results generalise easily to collections of independent particles. Choosing h as the unit of angular momentum, and regarding these commutation relations as the defining characteristic of angular momentum, the general quantal definition of angular momentum will be taken to be as follows: Angular momentum is a physical observable represented by three hermitian operators j x, j y and j z which satisfy the commutation relations [j x, j y ] = ij z, and cyclic permutations. These operators are the components of a vector j. The change in notation from L to j is intended to indicate the possibility of generalisation of the concept of angular momentum beyond that associated with classical orbital motion. The operators of angular momentum generate an algebra (the commutator of any two operators in the set is a linear combination of operators from the same set). Since the significance of operators in quantum mechanics lies in their matrix elements, there is obvious interest in establishing the matrix representations of the angular momentum algebra in terms of standard basis functions. Though no two components of the angular momentum operator commute with one another, all three components compute with the quadratic form j 2 = jx 2 + j2 y + j2 z, and it may be established that this is the most general angular momentum operator with this property. According to the general principles of quantum mechanics, j 2 may be diagonalised simultaneously with any one component of j, and their eigenvalues may be used to label quantum states. The standard choice is to diagonalise j 2 and j z, though any other choice of the component to be diagonalised would be completely equivalent. 2

3 It turns out to be useful to look for what may be thought of as eigenoperators of j 2 and j z, namely operators Θ which satisfy [ j 2, Θ] = λθ and [j z, Θ] = mθ, where λ and m are numbers. Any linear combination of the operators j α satisfies the first relation, with λ = 0, but the second is less trivial. It is easily established that the only linear combinations of j α satisfying the second eigenoperator relation are j z itself, trivially, with m = 0, and the two operators j ± = j x ± ij y, with m = ±1. (Of course, any multiples of these operators will satisfy the same relationship, since it is linear.) These operators satisfy the commutation relations [ j 2, j ± ] = 0 (1.2) [j z, j ± ] = ±j ± (1.3) [j +, j ] = 2j z (1.4) Consider a set of states λm which are simultaneous eigenstates of j 2 and j z with eigenvalues λ and m, respectively, i.e. j 2 λm = λ λm (1.5) j z λm = m λm (1.6) Applying to these states the commutation relations for j ± one obtains j 2 j ± λm = [ j 2, j ± ] λm + j ± j 2 λm = λj ± λm and j z j ± λm = [j z, j ± ] λm + j ± j z λm = (m ± 1)j ± λm from which it follows that j ± λm = α (±) λm λ, m ± 1 (1.7) where α (±) λm is a numerical coefficient and it is assumed that the states λm are all normalised. The operators j ± are therefore referred to as step operators, changing the eigenvalue of j z by one unit, up or down respectively, while leaving the eigenvalue of j 2 unchanged. It should be noted that j ± are not hermitian operators, but that each is the hermitian conjugate of the other. 3

4 and The relations j + j = j 2 x + j2 y + i[j y, j x ] = j 2 j 2 z + j z (1.8) j j + = j 2 j 2 z j z (1.9) which follow from the definition of j ±, may be used to determine the coefficients α (±) λm, since α (±) λm 2 = j ± λm 2 = λm j j ± λm = λm j 2 j 2 z j z λm = λ m 2 m. As always, quantal states are defined, even when normalised, only up to an overall phase. There is thus complete freedom to choose the phases of the basis states λm in order to simplify the above results as far as possible. The standard choice, which can be shown always to be possible, is known as the Condon-Shortley phase convention and involves choosing the coefficients α (±) λm to be real and positive. So α (±) λm = λ m 2 m. (1.10) Since the norm of a quantal state must be non-negative, it follows that 0 λm j +j + + j j λm = λm j j + + j + j λm = λm 2 j 2 2jz λm 2 = 2(λ m 2 ) so that m 2 λ and, for any finite value of λ, there is a limitation on the possible values of m. Consider a specific state λm and apply to it the angular-momentum raising (step-up) operator j +. This produces a state λ, m + 1. Repeated applications of j + will produce a succession of states λ, m + n, where n is an arbitrary positive integer. For any initial values of λ and m, there will eventually be some value of n for which (m + n) 2 > λ, which is forbidden. Thus there must exist some maximum value of n, denoted n max, such that 4

5 j + λ, m + n max = 0. Then j j + λ, m + n max = [λ (m + n max ) 2 (m + n max )] λ, m + n max = 0 i.e. λ = (m + n max )(m + n max + 1). Similarly, repeated applications of the step-down operator j to the state λm will ultimately produce a state λ, m n such that (m n ) 2 > λ, which is again forbidden. There must thus also exist some maximum value of the positive integer n, denoted n min, such that j λ, m n min = 0. Then j + j λ, m n min = [λ (m n min ) 2 + (m n min )] λ, m n min = 0 i.e. λ = (m n min )(m n min 1). Therefore (m + n max )(m + n max + 1) = (m n min )(m n min 1), from which m + n max = (m n min ) and so 2m = n min n max. Recalling that n max and n min are non-negative integers, it follows that m must be an integer or half an odd integer (positive or negative), with a maximum allowed value and a minimum allowed value for any fixed value of λ. Denoting the maximum allowed value of m by j, the above results imply that the minimum allowed value is j and that λ = j(j + 1). To summarise, it is inherent in the angular momentum algebra that the simultaneous eigenstates of j 2 and j z are determined by the quantum numbers j and m, where j is a non-negative number, either an integer or half an odd integer, and m takes the 2j +1 values j to j in integer steps. The normalised states jm satisfy j 2 jm = j(j + 1) jm (1.11) j z jm = m jm (1.12) j ± jm = α (±) jm j, m ± 1 (1.13) α (±) jm = j(j + 1) m(m ± 1) (1.14) where the last equation embodies a phase convention for the states. For each j there is a (2j + 1)-dimensional matrix representation for the three angular momentum operators j x, j y and j z, where the rows and columns are labelled by the quantum number m. Since j ± (and hence j x and j y ) connect every state to a neighbouring state, there is no smaller subset 5

6 of linear combinations of the states jm, for given j, which is closed under the operation of all three operators j α (i.e. such that any member of the subset is converted into a linear combination only of members of the subset by any of the operators j α ). Such representations are said to be irreducible. Since the whole of the above discussion has been based only on the commutation relations of the operators, it holds true for any set of operators satisfying the same commutator algebra and, in particular, for the orbital angular momentum operators l = ( r p)/ h = i r. It is well known that the simultaneous eigenfunctions of l 2 and l z are the spherical harmonics Y lm (θ, φ), where l is restricted to be an integer in order for the function to be singlevalued over the range of its arguments. The Condon-Shortley phase convention implies Y lm (θ, φ) = ( )m Y l, m (θ, φ). A very useful general result can now be established concerning the matrix elements, in the angular momentum basis, of any operator which commutes with all the components of the angular momentum operator j. Denoting the operator by Θ, the condition [Θ, j α ] = 0 for α = x, y, z implies also that [Θ, j 2 ] = 0. Matrix elements of this equation may be taken between states ξjm, where ξ symbolises all other quantum numbers required to specify the states completely. These quantum numbers are irrelevant to the present discussion and will simply remain unchanged at each step of the argument. Thus 0 = ξjm [Θ, j 2 ] ξ j m = ξjm Θ j 2 j 2 Θ ξ j m = j (j + 1) ξjm Θ ξ j m j(j + 1) ξjm Θ ξ j m = [j (j + 1) j(j + 1)] ξjm Θ ξ j m where the hermiticity of j 2 has been used in operating to the left on the state ξjm and in taking real eigenvalues. It follows that ξjm Θ ξ j m = 0 unless j = j. (1.15) In precisely the same way, taking matrix elements of [Θ, j z ] = 0 leads to ξjm Θ ξ j m = 0 unless m = m. (1.16) 6

7 Finally, since [Θ, j + ] = 0, it follows that ξj, m + 1 Θ ξ j, m + 1 = ξj, m + 1 Θj + ξ jm /α (+) jm = ξj, m + 1 j + Θ ξ jm /α (+) jm = α j,m+1 ξjm Θ ξ ( ) jm /α (+) jm = ξjm Θ ξ jm, from the explicit form of α (±) jm. Therefore, any operator which commutes with all the components of j is diagonal in the quantum numbers j and m and its matrix elements in the basis ξjm are independent of m. In particular, the unit operator commutes with all components of j. Its matrix elements simply express the overlap between different states. The general result here then ensures that the overlap between any two eigenstates of a given angular momentum operator j, no matter what other quantum numbers are required to specify the states completely, will be zero if the states belong to different eigenvalues j or m and will be independent of m when the states belong to the same eigenvalues j and m. II. COMBINATION OF ANGULAR MOMENTA Consider a number of independent angular momenta, i.e. a number of triplets of operators j (a), where a is an index labelling a particular triplet, such that the components of different operators j (a) commute, [j (a) α, j (b) β ] = iδ abj (a) γ (2.1) where (α, β, γ) is a cyclic permutation of (x, y, z). These may be the orbital angular momenta of different particles, or the orbital and spin angular momentum of a single particle, or some more general combination. Defining the resultant or total angular momentum operator j by j α = a j (a) α, (2.2) it is easily checked that the three components of j do in fact satisfy the angular momentum algebra, hence justifying the nomenclature. There will thus be a set of eigenstates jm with all the properties established above. 7

8 In this case, it is easy to find additional operators which commute with j 2 and j z and with one another and whose eigenvalues can be used to supply additional quantum numbers characterising the total angular momentum eigenstates. In fact, any operator ( j (a) ) 2 commutes with all the components of any j (b) and hence with j 2, j z and any ( j (b) ) 2. It is therefore possible to diagonalise simultaneously all the ( j (a) ) 2 operators, together with j 2 and j z, and to produce the states j 1 j 2 j 3...j n jm, with j 2 j 1 j 2 j 3...j n jm = j(j + 1) j 1 j 2 j 3...j n jm (2.3) j z j 1 j 2 j 3...j n jm = m j 1 j 2 j 3...j n jm (2.4) ( j (a) ) 2 j 1 j 2 j 3...j n jm = j a (j a + 1) j 1 j 2 j 3...j n jm. (2.5) No single component j (a) α of the individual angular momenta commutes with j 2 and j z, because of the cross terms in j 2. However, partial sums of angular momenta, like ( j (a) + j (b) ) 2 or ( j (a) + j (b) + j (c) ) 2, etc., do commute with all the ( j (a) ) 2 and with j 2 and j z, so many more angular momentum quantum numbers can be simultaneously specified. Some caution is required, though, since two partial sums with overlapping sets of indices a, b,... will not commute with one another unless one is wholly contained as a sub-sum in the other. Clearly, several different sets of mutually commuting operators can be found, all including j 2 and j z and each defining a different basis of eigenvectors. If the operator j 2 is omitted, many more bases can be defined, the simplest being that in which ( j (a) ) 2 and j z (a), for all a, are diagonal, namely j 1 m 1 j 2 m 2 j 3 m 3...j n m n. Others would involve various partial sums of angular momenta, such as j 1 j 2 j 3 j 123 m 123 j 4 j 5 j 45 m 45...j n m n, for example. There is an embarrassing richness of alternatives, not all of which are really different from one another (they may differ simply by a relabeling of the j (a), for instance), and not all of which are of practical importance. In practice, detailed calculations seldom deal explicitly with more than four angular momenta at one time, and then the number of really different bases is relatively small and the relations between them quite easily established. It is both simplest and most instructive to tackle, in order, the case of two, then three and then four angular momenta, after which the generalisation to larger numbers is straightforward. In the case of two independent angular momenta j (1) and j (2), there are only two complete sets of mutually commuting angular momentum operators, the first being 8

9 ( j (1) ) 2, j z (1), ( j (2) ) 2, j z (2) and the second being ( j (1) ) 2, ( j (2) ) 2, j 2, j z, where j = j (1) + j (2). The corresponding eigenvectors are j 1 m 1 j 2 m 2 and j 1 j 2 ; jm, the former being referred to as uncoupled, the latter as coupled. For given values of the quantum numbers j 1 and j 2, these two sets of eigenvectors constitute different orthonormal bases for the representation of the angular momentum operators and must be connected by a unitary transformation. (In accordance with the general result obtained at the end of the preceding section, eigenvectors corresponding to different values of the quantum number j 1 or j 2 are orthogonal to one another.) The transformation coefficients between bases are called Clebsch-Gordan coefficients or vector-coupling coefficients and are most easily written down in Dirac notation: j 1 m 1 j 2 m 2 = jm j 1 j 2 ; jm j 1 m 1 j 2 m 2 j 1 j 2 ; jm (2.6) j 1 j 2 ; jm = m 1 m 2 j 1 m 1 j 2 m 2 j 1 j 2 ; jm j 1 m 1 j 2 m 2 (2.7) where j 1 m 1 j 2 m 2 j 1 j 2 ; jm = j 1 j 2 ; jm j 1 m 1 j 2 m 2 and the unitarity conditions are j 1 j 2 ; j m j 1 j 2 m 1 m 2 j 1 j 2 m 1 m 2 j 1 j 2 ; jm = δ jj δ mm (2.8) m 1 m 2 j 1 m 1j 2 m 2 j 1 j 2 ; jm j 1 j 2 ; jm j 1 m 1 j 2 m 2 = δ m1 m δ 1 m 2 m. (2.9) 2 jm For given j 1 and j 2, the quantum numbers m 1 and m 2 have their standard ranges (from j i to j i in unit steps) and for given j, the quantum number m has its standard range, but it has not yet been determined what are the possible values of j. For this, it is necessary to look into some of the detailed properties of the vector coupling process. Intuitively, since j z = j z (1) + j z (2), it is reasonable to expect m = m 1 + m 2. Therefore, the largest possible value attainable by m is m max = j 1 + j 2. It follows that the largest value attainable by j is also j 1 + j 2, since if there were a larger value of j it would be associated with larger values of m. There is only one basis function j 1 m 1 j 2 m 2 with m 1 +m 2 = j 1 +j 2, so that, necessarily, j 1 j 2 ; j 1 +j 2, j 1 +j 2 = j 1 j 1 j 2 j 2 and the corresponding Clebsch-Gordan coefficient is unity. (Note the off-hand introduction of a further phase convention.) It can be checked by straightforward, if lengthy, calculation that the uncoupled state j 1 j 1 j 2 j 2 is an eigenstate of j 2 and of j z, with quantum numbers j 1 + j 2 and j 1 + j 2. The state j 1 j 2 ; j 1 +j 2, j 1 +j 2 1 is now produced by acting on this extreme state with the step-down operator j. Since the value m = j 1 +j 2 1 can be produced either by m 1 = j 1, m 2 = j 2 1 9

10 or by m 1 = j 1 1, m 2 = j 2, there are two independent linear combinations of uncoupled states with this m value. One of them will be produced by j j 1 j 2 ; j 1 + j 2, j 1 + j 2, and the orthogonal linear combination must then be the state j 1 j 2 ; j 1 + j 2 1, j 1 + j 2 1 (with once again a free choice of phase). This procedure may now be repeated m = j 1 +j 2 2 can be produced in three different ways: m 1 = j 1, m 2 = j 2 2; m 1 = j 1 1, m 2 = j 2 1; m 1 = j 1 2, m 2 = j 2 ; so there are three independent linear combinations of uncoupled states with this m value. Two of them are produced by operating with the step-down operator j on the two coupled states found in the preceding steps, and are j 1 j 2 ; j 1 +j 2, j 1 +j 2 2 and j 1 j 2 ; j 1 +j 2 1, j 1 +j 2 2. The third linear combination, orthogonal to both of these, is then necessarily j 1 j 2 ; j 1 +j 2 2, j 1 +j 2 2. Continuing in this way, the coupled states j 1 j 2 ; jm may be produced in stepwise fashion, each step automatically determining the corresponding Clebsch-Gordan coefficients (and requiring one further phase choice, which can obviously be used to make all the Clebsch- Gordan coefficients real, since the Condon-Shortley convention makes all the α (±) jm coefficients real). The above process produces one extra state at each step, with a j value decreased by unity from that of the previous step, and will terminate when all the uncoupled basis states have been used up, which happens when either m 1 or m 2 reaches the lower end of its range. At this point, the number of different ways of making the appropriate value of m is equal to the number of different j values produced in the preceding steps. At the end of the process, all (2j 1 + 1)(2j 2 + 1) uncoupled basis functions j 1 m 1 j 2 m 2 will have been used to produce an equal number of coupled basis functions j 1 j 2 ; jm : j 1 +j 2 (2j 1 + 1)(2j 2 + 1) = (2j + 1) j=j min = (j 1 + j 2 + 1) 2 j 2 min from which j min = j 1 j 2. Thus the two angular momenta j 1 and j 2 may be coupled to a total angular momentum j 1 j 2, j 1 j 2 + 1, j 1 j 2 + 2,...j 1 + j 2 1, j 1 + j 2, each j value occurring once. The same results may be obtained more formally in the following fashion. Applying j z to the defining equation j 1 j 2 ; jm = m 1 m 2 j 1 m 1 j 2 m 2 j 1 j 2 ; jm j 1 m 1 j 2 m 2 10

11 produces m j 1 j 2 ; jm = m 1 m 2 j 1 m 1 j 2 m 2 j 1 j 2 ; jm (m 1 + m 2 ) j 1 m 1 j 2 m 2. On substituting the defining equation again in the left hand side, this may be rearranged into the form m 1 m 2 (m m 1 m 2 ) j 1 m 1 j 2 m 2 j 1 j 2 ; jm j 1 m 1 j 2 m 2 = 0. The states being summed over are mutually orthogonal and hence linearly independent, so the sum can vanish only if the coefficient of every state in the sum vanishes, i.e. j 1 m 1 j 2 m 2 j 1 j 2 ; jm = 0 unless m = m 1 + m 2. Thus the additive condition on m assumed intuitively above arises as a selection rule on the Clebsch-Gordan coefficients. (Note that the state j 1 m 1 j 2 m 2 is an eigenstate of j z with eigenvalue m 1 + m 2, while j 1 j 2 ; jm is an eigenstate of j z with eigenvalue m. The Clebsch-Gordan coefficient is an overlap between these two states, so vanishes unless their j z eigenvalues are equal. This is an alternative proof of the m selection rule.) The same technique may be used to produce a pair of recursion relations for the Clebsch- Gordan coefficients, by applying the operators j ± to the defining equation, substituting the defining equation again on the left hand side of the result, rearranging and using the linear independence of the uncoupled basis functions. The properties of the coefficients α (±) jm and of the basis vectors j i m i at the extremes of the range of m are important in ensuring that the ranges of summation on the two sides of the equation are compatible. The result is α (±) jm j 1 m 1 j 2 m 2 j 1 j 2 ; j, m ± 1 = α (±) j 1,m 1 1 j 1, m 1 1, j 2 m 2 j 1 j 2 ; jm +α (±) j 2,m 2 1 j 1 m 1 j 2, m 2 1 j 1 j 2 ; jm. (2.10) These recursion relations can be used to express all the Clebsch-Gordan coefficients for given j 1 and j 2 in terms of either j 1 j 1 j 2, j j 1 j 1 j 2 ; jj or j 1, j j 2, j 2 j 2 j 1 j 2 ; jj. (Consider, for example, the case m = j, with the upper sign, for which the left hand side vanishes and the right hand side, with the aid of the m selection rule, produces a recursion relation in m 1 alone.) The magnitude of the single remaining undetermined coefficient can be found by applying the unitarity condition in the form m 1 m 2 j 1 m 1 j 2 m 2 j 1 j 2 ; jm 2 = 1, and its phase can be chosen arbitrarily. For convenience, the phase is chosen so that the relevant coefficient is real and positive, j 1 j 1 j 2, j j 1 j 1 j 2 ; jj > 0, and the recursion relation then 11

12 ensures that all the Clebsch-Gordan coefficients are real. Therefore, j 1 m 1 j 2 m 2 j 1 j 2 ; jm = j 1 j 2 ; jm j 1 m 1 j 2 m 2 and only one form of the coefficient will be used henceforth. Since the quantum numbers j 1 and j 2 remain the same on both sides of the coefficient, the abbreviated form j 1 m 1 j 2 m 2 jm is sufficient and will be adopted from now on. Given the limitations on the projection quantum numbers m, it is clearly necessary that j 2 j j 1 j 2, or else all the Clebsch-Gordan coefficients vanish. Since the whole exercise could equally well have been carried out in terms of j 1, j j 2, j 2 j 2 j 1 j 2 ; jj (or, for that matter, in terms of the two alternative forms with m 1 = j 1 and m 2 = j 2 ), it is also necessary that j 1 j j 2 j 1. Combining the two sets of restrictions, it is seen that the values of j must satisfy the triangle inequality j 1 j 2 j j 1 + j 2. (2.11) From the general requirement that j be an integer or half an odd integer, and that m have the same character as j, and from the selection rule m = m 1 +m 2, it follows that j will be an integer if j 1 and j 2 are both integers or both half odd integers, and will be half an odd integer if only one of j 1 and j 2 is an integer. It is also worth noting that the whole treatment of the coupling of j 1 and j 2 was completely symmetric between the two component angular momenta until the choice of the phase of the determining coefficient in the recursion relation. The choice j 1 j 1 j 2, j j 1 j 1 j 2 ; jj > 0, rather than j 1, j j 2, j 2 j 2 j 1 j 2 ; jj > 0, introduces an asymmetry between j 1 and j 2, so that the coefficients j 1 m 1 j 2 m 2 jm and j 2 m 2 j 1 m 1 jm can differ at most by a sign. Similarly, since the recursion relation could have been used to relate all Clebsch-Gordan coefficients to those with m = j, instead of m = j, there will be a possible sign difference between coefficients with m 1, m 2, m or with m 1, m 2, m. Detailed discussion of the recursion relation, or inspection of the complicated explicit formula for the Clebsch-Gordan coefficients derived from group theory by various authors and quoted in the references, allow the extraction of the following symmetry properties of the coefficients: j 1 m 1 j 2 m 2 jm = ( ) j 1+j 2 j j 2 m 2 j 1 m 1 jm (2.12) = ( ) j 1+j 2 j j 1, m 1, j 2, m 2 j, m (2.13) = ( ) j 2+m 2 (2j + 1)/(2j 1 + 1) j 2, m 2 jm j 1 m 1 (2.14) = ( ) j 1 m 1 (2j + 1)/(2j 2 + 1) jmj 1, m 1 j 2 m 2. (2.15) 12

13 These properties suggest a basic symmetry among the three angular momenta involved the two constituent angular momenta j 1 and j 2 and the total angular momentum j. The same underlying symmetry is indicated by the triangle inequality satisfied by j 1, j 2 and j, which is equally valid for any division of the three angular momenta into two constitutents and a resultant. This symmetry is effectively exploited in an alternative form of the vectorcoupling coefficients, the Wigner 3-j symbol, defined by j 1 j 2 j 3 = ( ) j 1 j 2 m 3 j 1 m 1 j 2 m 2 j 3, m 3 / m 1 m 2 m 3 2j (2.16) The 3-j symbol vanishes unless j 1, j 2, j 3 satisfy the triangle inequality and m 1 +m 2 +m 3 = 0. It is left unchanged in value by any cyclic permutation of its three columns, but is multiplied by the phase factor ( ) j 1+j 2 +j 3 if any two columns are interchanged or if the signs of all three projection quantum numbers m i are reversed. Thus j 1 j 2 j 3 = 0 unless j 1 + j 2 + j 3 is even. The symbol also has the useful property that j 1 j 2 j 3 j 1 m 1 j 2 m 2 j 3 m 3 m 1 m 2 m 3 m 1 m 2 m 3 represents a state of three angular momenta coupled to total angular momentum zero which is symmetric under cyclic permutation of the three angular momenta. Some consequences of the angular-momentum coupling formalism are of general interest. In the case j = 0, the recursion relation for the Clebsch-Gordan coefficients becomes 0 = α (+) j 1,m 1 1 j 1, m 1 1, j 2, m α (+) j 2 m 2 j 1 m 1 j 2 m From the selection rules on the coefficients, necessarily j 1 = j 2 and m 1 = m 2. But from the explicit expression for the α (±) jm coefficients, α (+) j,m 1 = α (+) j, m, so the result reduces to jmj, m 00 = j, m 1, j, (m 1) 00, from which it follows that jmj, m 00 = ( ) j m jjj, j 00. But m jmj, m 00 2 = 1, so that jjj, j 00 = 1/ 2j + 1, where the standard phase convention has been used to select the positive square root. Finally, j 1 m 1 j 2 m 2 00 = ( ) j 1 m 1 δ j1 j 2 δ m1, m 2 / 2j (2.17) Using the symmetry relation of the Clebsch-Gordan coefficient, this may be rewritten in the form j 1 m 1 00 j 2 m 2 = δ j1 j 2 δ m1 m 2, which is self-evident. Alternatively, j 1 j 2 0 = ( ) j 1 m 1 δ j1 j 2 δ m1, m 2 / 2j (2.18) m 1 m

14 Another special case of the Clebsch-Gordan coefficient which is frequently useful can be obtained straightforwardly from the explicit formula for the coefficient, alluded to previously, namely jjj, j J0 = (2j)! (2J + 1)/(2j + J + 1)!(2j J)!. (2.19) Two identical systems, defined by their angular momentum j and by additional quantum numbers collectively denoted by ξ, can be coupled to a state of well-defined total angular momentum J as follows: (ξj) 2 JM = m 1 m 2 jm 1 jm 2 JM ξjm 1 ξjm 2. The dummy indices m i have the same range and may be exchanged to produce (ξj) 2 JM = m 1 m 2 jm 2 jm 1 JM ξjm 2 ξjm 1 = ( ) j+j J m 1 m 2 jm 1 jm 2 JM ξjm 2 ξjm 1, by the symmetry property of the Clebsch-Gordan coefficient. This may be rewritten (ξj) 2 JM = ( ) 2j J P 12 m 1 m 2 jm 1 jm 2 JM ξjm 1 ξjm 2 where P 12 is a permutation operator which exchanges the two systems in any state on which it acts (but has no effect on numerical coefficients like the Clebsch-Gordan coefficients). This final result states that P 12 (ξj) 2 JM = ( ) 2j J (ξj) 2 JM, (2.20) so that the coupled state (ξj) 2 JM is automatically symmetric or antisymmetric under interchange of the two systems, according as 2j J is even or odd, respectively. (Note the importance of the identity of the systems. For ξ 1 ξ 2, this result would not hold.) Considering that identical bosons, which have integer angular momentum, are required to form totally symmetric states, while identical fermions, which have half-odd-integer angular momentum, are required to form totally antisymmetric states, it follows that a pair of identical fermions or a pair of identical bosons can couple only to even total angular momentum J. This argument can be extended, in the case where each system is characterised by two or more independent angular momenta, to produce a correspondence between the sign of the phase factor ( ) J 1+J 2 + +J n and the symmetry of the state (ξj 1 j 2...j n ) 2 J 1 M 1 J 2 M 2...J n M n. Thus for systems characterised by angular momentum 14

15 and isospin, for example, selection rules arise for J + T, where T is the total isospin of the pair of identical systems. III. RECOUPLING OF ANGULAR MOMENTA As indicated at the beginning of the previous section, there will generally be more than two different bases applicable to problems involving more than two angular momenta. In the case of three angular momenta, the bases of interest are the totally uncoupled basis j 1 m 1 j 2 m 2 j 3 m 3 and the totally coupled basis j 1 j 2 j 3 jm, where the latter is not yet completely specified, being determined by the eigenvalues of only five, rather than six, mutually commuting operators. The totally coupled basis may be fully specified by defining one intermediate coupling, which can be done in three different ways (j 1 j 2 )j 12 j 3 jm, (j 1 j 3 )j 13 j 2 jm or j 1 (j 2 j 3 )j 23 jm, where the notation should be self-explanatory. The transition from the fully uncoupled to the fully coupled bases goes through the intermediate partially coupled basis, with Clebsch-Gordan coefficients being used to couple pairs of angular momenta at each step. So (j 1 j 2 )j 12 j 3 jm = j 12 m 12 j 3 m 3 j 12 m (j 1 j 2 )j 12 m 12 j 3 m 3 (3.1) m 12 m 3 = j 12 m 12 j 3 m 3 jm j 1 m 1 j 2 m 2 j 12 m 12 j 1 m 1 j 2 m 2 j 3 m 3, (3.2) m 1 m 2 m 3 m 12 and similarly for each of the other possible intermediate couplings. However, the existence of different bases does not imply a proliferation of states any one of the three fully coupled bases is equally valid, so there must exist unitary transformations between them. The vectors belonging to a particular basis, such as (j 1 j 2 )j 12 j 3 jm, are distinguished by the values of the six quantum numbers listed and are all mutually orthogonal, according to the general result proved at the end of section 1. However, vectors from different bases are not necessarily orthogonal, provided they have the same values of the common quantum numbers j 1, j 2, j 3, j and m. The overlaps between functions belonging to different bases are the coefficients of the unitary transformation between the bases, as is clear in Dirac notation: (j 1 j 2 )j 12 j 3 jm = j 1 (j 2 j 3 )j 23 (j 1 j 2 )j 12 j 3 j j 1 (j 2 j 3 )j 23 jm. (3.3) j 23 The transformation coefficient, again by the general result proved at the end of section 1, 15

16 is diagonal in j and m and independent of the value of m, which is thus dropped from the symbol for the overlap. Using the decomposition of the totally coupled states into totally uncoupled states in terms of Clebsch-Gordan coefficients, the transformation coefficient is explicitly written as j 1 (j 2 j 3 )j 23 j (j 1 j 2 )j 12 j 3 j = m 1 m 2 m 3 m 12 m 23 j 1 m 1 j 2 m 2 j 12 m 12 j 12 m 12 j 3 m 3 jm j 1 m 1 j 23 m 23 jm j 2 m 2 j 3 m 3 j 23 m 23. (3.4) Since the Clebsch-Gordan coefficients are real in the standard phase convention, so are these transformation (or recoupling) coefficients. Note that the recoupling coefficients are determined purely by the Clebsch-Gordan coefficients and are independent of any property of the states considered except their angular momentum quantum numbers. The recoupling coefficient is commonly written in terms of a more symmetrical quantity, the Wigner 6-j symbol, as follows: (j 1 j 2 )j 12 j 3 j j 1 (j 2 j 3 )j 23 j = ( ) j 1+j 2 +j 3 j +j 1 j 2 j 12 (2j )(2j ) j 3 j j 23. (3.5) An alternative notation is in terms of the Racah coefficient W(j 1 j 2 jj 3 ; j 12 j 23 ) = ( ) j 1+j 2 +j 3 j +j 1 j 2 j 12 j 3 j j 23. (3.6) The 6-j symbol is real and its selection rules are the four triangle inequalities, (j 1 j 2 j 12 ), (j 12 j 3 j), (j 2 j 3 j 23 ) and (j 1 j 23 j), where (j 1 j 2 j 3 ) j 1 j 2 j 3 j 1 + j 2. (3.7) Its symmetry properties may be read off from those of the Clebsch-Gordan coefficients (which is much more easily done when the latter are replaced by 3-j symbols). It turns out to be invariant under those permutations of its six arguments which leave the set of four triangle inequalities unchanged. Depicting these inequalities in the self-explanatory form it is quickly seen that this set of four pictures, and hence the 6-j symbol, is unchanged under arbitrary permutation of its three columns or under simultaneous inversion a b b a of any 16

17 pair of columns. Since the 6-j symbol is entirely independent of any projection quantum numbers m, it may be regarded as a rotational invariant. Other transformation coefficients can be expressed in terms of the one explicitly investigated above by suitable reorderings of the angular momenta being coupled, using the symmetries of the Clebsch-Gordan coefficients. For instance, (j 1 j 3 )j 13 j 2 j (j 1 j 2 )j 12 j 3 j = ( ) j 13+j 2 j j 2 (j 1 j 3 )j 13 j (j 1 j 2 )j 12 j = ( ) j 13+j 2 j ( ) j 1+j 2 j 12 j 2 (j 1 j 3 )j 13 j (j 2 j 1 )j 12 j 3 j = ( ) j 12+j 13 j 1 j ( ) j 2+j 1 +j 3 +j j 2 j 1 j 12 (2j )(2j ) j 3 j j 13 = ( ) j 2+j 3 +j 12 +j 13 j 2 j 1 j 12 (2j )(2j ) j 3 j j 13. (It should be noted that the 6-j symbol is defined by an overlap in which the three constituent angular momenta appear in precisely the same order in both totally coupled states. The reality of the 6-j symbol makes the order of the two totally coupled states in the overlap unimportant.) In manipulating phase factors, as has been done above, the following observations are useful. When j 1 and j 2 are coupled to produce j 3, any of the three may be, in principle, an integer or a half-odd-integer. However, as remarked above, in all cases either all three or only one of the j s must be integral. This is sufficient to ensure that the combinations j 1 + j 2 ± j 3 are integers, so that ( ) j 1+j 2 j 3, for instance, is real and is equal to its inverse, ( ) j 1 j 2 +j 3. However, ( ) j 1+j 2 j 3 is not necessarily equal to ( ) j 1+j 2 +j 3 these two phases are equal if j 3 is an integer, but opposite if it is a half-odd-integer. Similar arguments hold for j 1, j 2 and j 3 coupled to produce j, where j 1 + j 2 + j 3 + j is always an integer and ( ) j 1+j 2 +j 3 +j = ( ) j 1 j 2 j 3 j is real. Also, for any angular momentum j and its projection m, ( ) 4j = 1 = ( ) 4m and ( ) j m = ( ) m j is real, as is ( ) j+m = ( ) j m. From the unitarity of the transformation between bases and the definition of the 6-j symbol, the following useful relationships may be written down. j 1 j 2 j 12 j 1 j 2 j 12 (2j ) j 23 j 3 j j 23 j 3 j j 23 = δ j 12 j 12 /(2j ) (3.8) 17

18 j 1 m 1 j 2 m 2 j 12 m 12 j 12 m 12 j 3 m 3 jm = ( ) j 1+j 2+j 3+j j 1 j 2 j 12 (2j )(2j ) m 12 j 23 m 23 j 3 j j 23 j 1 m 1 j 23 m 23 jm j 2 m 2 j 3 m 3 j 23 m 23 (3.9) j 1 m 1 j 2 m 2 j 12 m 12 j 12 m 12 j 3 m 3 jm j 2 m 2 j 3 m 3 j 23 m 23 m 2 m 3 m 12 = ( ) j 1+j 2 +j 3 j +j 1 j 2 j 12 (2j )(2j ) j 3 j j 23 j 1m 1 j 23 m 23 jm (3.10) where the second equation is simply a rewriting of the definition of the unitary transformation coefficient and the third is derived from it by use of the orthogonality of the Clebsch-Gordan coefficients. A sum rule for the 6-j symbol may be derived by considering different ways of achieving the same ultimate recoupling of three angular momenta. Clearly, (j 1 j 2 )j 12 j 3 j j 2 (j 3 j 1 )j 13 j = (j 1 j 2 )j 12 j 3 j j 1 (j 2 j 3 )j 23 j j 1 (j 2 j 3 )j 23 j j 2 (j 3 j 1 )j 13 j. j 23 The transformation coefficients on both sides can be reduced to standard form by changing the order of coupling of pairs of angular momenta, with the appropriate phase factors, and rewritten in terms of 6-j symbols to yield ( ) j 12+j 13+j 23 j 1 j 2 j 12 j 1 j 3 j 13 (2j ) j 23 j 3 j j 23 j 2 j j 23 = j 1 j 3 j 13 j j 2 j 12. (3.11) A somewhat more complex sum rule is obtained by considering the recoupling of four angular momenta, three at a time, in different ways. In an obvious notation, (j 1 j 2 )j 12 j 3 j 123 j 4 j (j 2 j 3 )j 23 j 1 j 123 j 4 j (j 2 j 3 )j 23 (j 1 j 4 )j 14 j; or (j 1 j 2 )j 12 j 3 j 123 j 4 j (j 1 j 2 )j 12 j 4 j 124 j 3 j (j 1 j 4 )j 14 j 2 j 124 j 3 j (j 1 j 4 )j 14 (j 2 j 3 )j 23 j. The resulting equality can be converted into the Biedenharn-Elliott sum rule for 6-j symbols, ( ) j 1+j 2+j 3+j 4+j 12+j 23+j 14+j 123+j j 124+j 1 j 2 j 12 j 12 j 3 j 123 (2j ) j 124 j 124 j 4 j 14 j j 4 j 124 j 2 j 3 j 23 j j 14 j 124 = j 1 j 23 j 123 j 1 j 23 j 123 j j 4 j 14 j 3 j 12 j 2. (3.12) 18

19 The value of the 6-j symbol in the special case where one of its arguments is zero is easily determined by substitution in the equation relating the 6-j symbol to the Clebsch-Gordan coefficients, or by observing that (j 1 j 2 )j 12 0j j 1 (j 2 0)j 23 j = δ j12 jδ j2 j 23. Either procedure produces the result j 1 j 2 j j 2 j 1 0 = ( )j 1+j 2 +j δ j1 j 1 δ j 2 j 2 / (2j 1 + 1)(2j 2 + 1). (3.13) The symmetry properties of the 6-j symbol can then be used to evaluate a 6-j symbol any of whose arguments is zero. Inserting this special value of the 6-j symbol into the orthogonality relation or any of the sum rules for these symbols, by setting one of the free angular momenta (not an index of summation) equal to zero, will generally produce a trivial identity, but in a few cases it produces further useful sum rules. Setting j 12 = 0 in the orthogonality relation leads to ( ) j 1+j 2+j 12 j 1 j 2 j 12 (2j ) j 12 j 2 j 1 j = (2j 1 + 1)(2j 2 + 1)δ j0. (3.14) Similarly, setting j 13 = 0 in the first sum rule above produces j 1 j 2 j 12 (2j + 1) j j 1 j 2 j = ( )2(j 1+j 2 ), which can be rewritten in the form j ( ) 2j 1 j 2 j 12 (2j + 1) = 1. (3.15) j j 1 j 2 j Setting an argument equal to zero in the Biedenharn-Elliott sum rule produces no new results. For the case of four coupled angular momenta, the uncoupled basis j 1 m 1 j 2 m 2 j 3 m 3 j 4 m 4 is related to the fully coupled basis j 1 j 2 j 3 j 4 j j jm (where j and j are intermediate couplings required to complete a full set of eight mutually commuting operators) by a unitary transformation. The number of possible sets of fully coupled basis functions is now quite 19

20 large (illustrative examples are j = j 12, j = j 123 ; j = j 13, j = j 134 ; j = j 12, j = j 34 ), but they fall into two classes bases where two angular momenta are coupled together, the result coupled to a third and that result to the fourth angular momentum, or bases where two distinct pairs of angular momenta are each coupled together and then their resultants are coupled to produce the total angular momentum. Any two such bases are related by a unitary transformation. The first type is of considerably less interest than the second and can be handled by successive uses of the 6-j symbol. The second type is sufficiently useful to have warranted the introduction of a new recoupling coefficient. The prototype transformation coefficient for four angular momenta coupled pairwise is (j 1 j 2 )j 12 (j 3 j 4 )j 34 j (j 1 j 3 )j 13 (j 2 j 4 )j 24 j, where again the overlap between states from different bases is diagonal in j and m and independent of m, which is dropped from the transformation bracket. (The overlap is also, of course, diagonal in the common quantum numbers j 1, j 2, j 3, j 4.) This can be written in terms of Clebsch-Gordan coefficients as j 1 m 1 j 2 m 2 j 12 m 12 j 3 m 3 j 4 m 4 j 34 m 34 j 12 m 12 j 34 m 34 jm m 1 m 2 m 3 m 4 m 12 m 34 m 13 m 24 j 1 m 1 j 3 m 3 j 13 m 13 j 2 m 2 j 4 m 4 j 24 m 24 j 13 m 13 j 24 m 24 jm, and is hence real. Once again it turns out to be convenient to introduce a slightly different recoupling coefficient, the Wigner 9-j symbol, defined by (j 1 j 2 )j 12 (j 3 j 4 )j 34 j (j 1 j 3 )j 13 (j 2 j 4 )j 24 j = j 1 j 2 j 12 (2j )(2j )(2j )(2j ) j 3 j 4 j 34. (3.16) j 13 j 24 j Converting the Clebsch-Gordan coefficients to 3-j symbols and exploiting the m- independence of the recoupling coefficient, it is found that j 1 j 2 j 12 j 3 j 4 j 34 = j 1 j 2 j 12 j 3 j 4 j 34 j 13 j 24 j j 13 j 24 j all m s m 1 m 2 m 12 m 3 m 4 m 34 m 13 m 24 m j 1 j 3 j 13 j 2 j 4 j 24 j 12 j 34 j (3.17) m 1 m 3 m 13 m 2 m 4 m 24 m 12 m 34 m where each row and each column of the 9-j symbol is represented by a 3-j symbol and all 20

21 nine projection quantum numbers m are summed over. The result, independent of projection quantum numbers, may be regarded as a rotational invariant. The selection rules for the 9-j symbol are now trivial each row and each column of the symbol contributes a triangle inequality which must be satisfied if the symbol is not to vanish. The symmetries of the 9-j symbol are also easily read off: the interchange of a pair of rows or of a pair of columns produces a phase factor ( ) j 1+j 2 +j 3 +j 4 +j 12 +j 34 +j 13 +j 24 +j, while transposition of the symbol (writing the rows as columns and the columns as rows by reflecting in the main diagonal) simply reorders the six 3-j symbols inside the summation and so leaves the symbol unchanged in value. The former symmetry introduces a further selection rule if any two columns or any two rows of the symbol are identical, then it vanishes unless the sum of its nine arguments is even (or, equivalently, unless the sum of the angular momenta in the remaining column or row is even). As before, the unitarity of the transformation imposes orthogonality conditions on the 9-j symbols, in the form j 1 j 2 j 12 j 1 j 2 j 12 (2j )(2j ) j 3 j 4 j 34 j 3 j 4 j = δ 34 j12 j δ 12 j 34 j /(2j )(2j ). j 13 j 24 j 13 j 24 j j 13 j 24 j (3.18) The 9-j symbol can be expressed in terms of 6-j symbols by recoupling four angular momenta in stages, with three angular momenta being recoupled at each stage, as follows: (j 1 j 2 )j 12 (j 3 j 4 )j 34 j (j 1 j 3 )j 13 (j 2 j 4 )j 24 j = j 123 (j 1 j 2 )j 12 (j 3 j 4 )j 34 j (j 1 j 2 )j 12 j 3 j 123 j 4 j (j 1 j 2 )j 12 j 3 j 123 j 4 j (j 1 j 3 )j 13 j 2 j 123 j 4 j (j 1 j 3 )j 13 j 2 j 123 j 4 j (j 1 j 3 )j 13 (j 2 j 4 )j 24 j from which it follows that j 1 j 2 j 12 j 3 j 4 j 34 = j ( ) 2j j 2 j 12 j 3 j 4 j 34 j 13 j 24 j (2j ) j 123 j j 13 j 24 j 123 j 3 j 13 j j 12 j 123 j 4 j 123 j 2. (3.19) As in the case of the expression for the 6-j symbol in terms of Clebsch-Gordan coefficients, this result may be manipulated, using the orthogonality of the 6-j symbols, to express a product of two 6-j symbols as a sum of products of a 6-j and a 9-j symbol. Successive recouplings of four angular momenta can be used to obtain a sum rule expressing a single 21

22 9-j symbol as a sum of products of pairs of 9-j symbols. The value of the 9-j symbol when one of its arguments vanishes is also easily obtained from the expression in terms of 6-j symbols. Given the symmetries of the 9-j symbol, it is sufficient to evaluate j 1 j 2 j 12 j 3 j 4 j 34 = ( ) j 2+j 12 +j 3 +j 13 j 1 j 2 j 12 δ j12 j 34 δ j13 j 24 j j 13 j j 3 j 13 / (2j )(2j ). (3.20) A useful special case arises when a whole row or column of arguments vanishes. This is covered by the specific result j 1 j 1 0 j 2 j 2 0 = δ j1 j δ 1 j 2 j δ 2 j 3 j (2j / )(2j 2 + 1)(2j 3 + 1), (3.21) j 3 j 3 0 where j 1, j 2 and j 3 satisfy the triangle relation. IV. SPHERICAL TENSOR OPERATORS Now that the quantum states of a system have been characterized in terms of their angular momentum properties, it becomes of interest to investigate the effect on these properties of various operators which act in the space of states. Consider, in the simplest case, the position and momentum operators of a single spinless particle with orbital angular momentum operator L = ( r p)/ h. Using the standard commutation relation [r α, p β ] = i hδ αβ, where the indices α, β represent the cartesian components x, y, z, it is easily established that the commutator [L α, r β ] is a simple linear combination of the components of r, while [L α, p β ] is the same linear combination of the corresponding components of p. It is possible to define specific linear combinations r m and p m of the components of r and p respectively, with m = 1, 0, +1, such that where v represents either r or p and where [L z, v m ] = mv m (4.1) [L ±, v m ] = α (±) 1m v m±1 (4.2) v 0 = v z (4.3) v ±1 = 1 (v x ± iv y ). 2 (4.4) 22

23 These eigenoperators of L have properties reminiscent of the defining properties of a set of states of angular momentum 1, projection m, with the commutator of L with the operators v m playing the role of the action of the operator L on the states 1m. This property of the operators r and p can be generalized and is extremely useful in all applications of the angular momentum algebra. Before going on to the general case, consider the set of operators r α p β, the nine possible products of the cartesian components (α, β = x, y, z) of r and p. Once again, the commutators of L with these operators produce only linear combinations of operators in the same set. This is guaranteed by the fact that r and p have this property, while commutators satisfy the identity [A, BC] = [A, B]C + B[A, C]. The operator products r α p β fall naturally into three subsets the combination r p (the sum of the diagonal elements of the set, called the trace), the vector product r p (the three antisymmetric combinations r α p β r β p α ) and the five independent symmetric combinations r α p β + r β p α 2 3 δ αβ r p, with vanishing trace. Each product r α p β can be written as a sum of terms from the three subsets. It is then found that each of these subsets is closed under commutation with L, i.e. the commutator of any component of L with a member of one of the subsets is a linear combination of members of the same subset. The scalar product r p commutes with all the components of L; the vector product r p has the same commutation relations with L as the individual vectors r and p; while specific linear combinations ( r p) m of the elements of the symmetric subset can be found for which [L z, ( r p) m ] = m( r p) m (4.5) [L ±, ( r p) m ] = α (±) 2m ( r p) m±1. (4.6) These results generalize to other operators than r, p and their products and to more general angular momentum operators than L. Given any set of operators which is closed under commutation with the components of the angular momentum j, it can always be broken up into subsets, each of which is itself closed under commutation with j but cannot be broken up into smaller subsets with the same property of closure. Any of the original operators can be written as a linear combination of operators from these subsets. Within each such irreducible subset, specific linear combinations can be chosen so that their commutation relations with j have the standard form encountered twice above, analogous to 23

24 that of sets of states jm acted upon by the operator j. Further it can be demonstrated that any arbitrary operator can be written as a linear combination of operators belonging to such standard sets. These general statements are systematized in the definition of an (irreducible) spherical tensor operator. An (irreducible) spherical tensor operator T (j) m is one of a set of 2j + 1 operators, corresponding to different values of m = j, j + 1,..., j 1, j, which satisfy The spherical tensor operator T (j) m [j z, T m (j) (j) ] = mt m (4.7) [ ] j±, T m (j) (±) = α jm T m±1. (j) (4.8) is said to be of rank j, projection m. This definition essentially generalizes the notion of angular momentum of states to include that of angular momentum of operators. It is now possible to investigate the effect on the angular momentum of a state of the action of a spherical tensor operator. Consider the state produced by operating with T (j) m on j m. It is still an eigenstate of j z, since j z T (j) m j m = [j z, T (j) m ] j m + T (j) m j z j m = mt (j) m j m + T (j) m m j m = (m + m )T (j) m j m, but its eigenvalue has become m + m. It is tedious but straightforward to check that T (j) m j m is not, however, an eigenstate of j 2. But it is essentially a product of two objects of well-defined angular momenta j and j, with well-defined projection quantum numbers m and m. The analogy with the product state j 1 m 1 j 2 m 2 suggests that an appropriate linear combination of such products using, of course, the Clebsch-Gordan coefficients may have well-defined angular momentum. Consider, therefore, the state JM jj = mm jmj m JM T (j) m j m. Because of the Clebsch-Gordan selection rule M = m + m, this is, in fact, an eigenstate of 24

25 j z with eigenvalue M, as suggested by the notation. Now, j ± JM jj = mm jmj m JM {[j ±, T (j) m ] + T (j) m j ± } j m = jmj m JM {α (±) jm T m±1 j (j) m + α (±) mm (j) j m T m j, m ± 1 } = mm [α (±) j,m 1 j, m 1, j m JM + α (±) j,m 1 jmj, m 1 JM ]T (j) m j m = mm α (±) JM jmj m J, M ± 1 T (j) m j m = α (±) JM J, M ± 1 jj where use has been made of the Clebsch-Gordan coefficient recursion relations and of the restrictions on the ranges of the projection quantum numbers. But this result is precisely what is required to establish JM jj as an eigenstate of j 2 with eigenvalue J(J + 1), thus fully justifying the notation used. The Clebsch-Gordan coefficients thus couple the rank of a spherical tensor operator to the angular momentum of the state on which it acts so that the state so produced is a welldefined angular momentum eigenstate. The orthogonality of the Clebsch-Gordan coefficients can now be used to write T (j) m j m = JM jmj m JM JM jj, which can be used to evaluate the matrix element j m T (j) m j m = JM jmj m JM j m JM jj = JM jmj m JM Θ j jj δ j Jδ m M = jmj m j m Θ j jj, where Θ j jj, the overlap between the states j m and JM jj, is diagonal in j (J) and in m (M) and independent of M. Therefore the dependence of the matrix element on the projection quantum numbers is contained entirely in the Clebsch-Gordan coefficient and is independent of the detailed dynamics of the states and of the operator involved. This very powerful and extremely useful result is known as the Wigner-Eckart theorem. It should be emphasized that all that is required for the theorem to hold is for the two states concerned to be eigenstates of the same angular momentum operator relative to which the operator involved is a spherical tensor operator. Any such matrix element can then be regarded as 25

State of Stress at Point

State of Stress at Point State of Stress at Point Einstein Notation The basic idea of Einstein notation is that a covector and a vector can form a scalar: This is typically written as an explicit sum: According to this convention,

More information

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS Systems of Equations and Matrices Representation of a linear system The general system of m equations in n unknowns can be written a x + a 2 x 2 + + a n x n b a

More information

MATH10212 Linear Algebra. Systems of Linear Equations. Definition. An n-dimensional vector is a row or a column of n numbers (or letters): a 1.

MATH10212 Linear Algebra. Systems of Linear Equations. Definition. An n-dimensional vector is a row or a column of n numbers (or letters): a 1. MATH10212 Linear Algebra Textbook: D. Poole, Linear Algebra: A Modern Introduction. Thompson, 2006. ISBN 0-534-40596-7. Systems of Linear Equations Definition. An n-dimensional vector is a row or a column

More information

Particle Physics. Michaelmas Term 2011 Prof Mark Thomson. Handout 7 : Symmetries and the Quark Model. Introduction/Aims

Particle Physics. Michaelmas Term 2011 Prof Mark Thomson. Handout 7 : Symmetries and the Quark Model. Introduction/Aims Particle Physics Michaelmas Term 2011 Prof Mark Thomson Handout 7 : Symmetries and the Quark Model Prof. M.A. Thomson Michaelmas 2011 206 Introduction/Aims Symmetries play a central role in particle physics;

More information

A Primer on Index Notation

A Primer on Index Notation A Primer on John Crimaldi August 28, 2006 1. Index versus Index notation (a.k.a. Cartesian notation) is a powerful tool for manipulating multidimensional equations. However, there are times when the more

More information

The Matrix Elements of a 3 3 Orthogonal Matrix Revisited

The Matrix Elements of a 3 3 Orthogonal Matrix Revisited Physics 116A Winter 2011 The Matrix Elements of a 3 3 Orthogonal Matrix Revisited 1. Introduction In a class handout entitled, Three-Dimensional Proper and Improper Rotation Matrices, I provided a derivation

More information

Linear Algebra Notes for Marsden and Tromba Vector Calculus

Linear Algebra Notes for Marsden and Tromba Vector Calculus Linear Algebra Notes for Marsden and Tromba Vector Calculus n-dimensional Euclidean Space and Matrices Definition of n space As was learned in Math b, a point in Euclidean three space can be thought of

More information

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 2. x n. a 11 a 12 a 1n b 1 a 21 a 22 a 2n b 2 a 31 a 32 a 3n b 3. a m1 a m2 a mn b m

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 2. x n. a 11 a 12 a 1n b 1 a 21 a 22 a 2n b 2 a 31 a 32 a 3n b 3. a m1 a m2 a mn b m MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS 1. SYSTEMS OF EQUATIONS AND MATRICES 1.1. Representation of a linear system. The general system of m equations in n unknowns can be written a 11 x 1 + a 12 x 2 +

More information

Chapter 9 Unitary Groups and SU(N)

Chapter 9 Unitary Groups and SU(N) Chapter 9 Unitary Groups and SU(N) The irreducible representations of SO(3) are appropriate for describing the degeneracies of states of quantum mechanical systems which have rotational symmetry in three

More information

Notes on Determinant

Notes on Determinant ENGG2012B Advanced Engineering Mathematics Notes on Determinant Lecturer: Kenneth Shum Lecture 9-18/02/2013 The determinant of a system of linear equations determines whether the solution is unique, without

More information

Continued Fractions and the Euclidean Algorithm

Continued Fractions and the Euclidean Algorithm Continued Fractions and the Euclidean Algorithm Lecture notes prepared for MATH 326, Spring 997 Department of Mathematics and Statistics University at Albany William F Hammond Table of Contents Introduction

More information

Introduction to Matrix Algebra

Introduction to Matrix Algebra Psychology 7291: Multivariate Statistics (Carey) 8/27/98 Matrix Algebra - 1 Introduction to Matrix Algebra Definitions: A matrix is a collection of numbers ordered by rows and columns. It is customary

More information

Similarity and Diagonalization. Similar Matrices

Similarity and Diagonalization. Similar Matrices MATH022 Linear Algebra Brief lecture notes 48 Similarity and Diagonalization Similar Matrices Let A and B be n n matrices. We say that A is similar to B if there is an invertible n n matrix P such that

More information

Numerical Analysis Lecture Notes

Numerical Analysis Lecture Notes Numerical Analysis Lecture Notes Peter J. Olver 5. Inner Products and Norms The norm of a vector is a measure of its size. Besides the familiar Euclidean norm based on the dot product, there are a number

More information

Orthogonal Diagonalization of Symmetric Matrices

Orthogonal Diagonalization of Symmetric Matrices MATH10212 Linear Algebra Brief lecture notes 57 Gram Schmidt Process enables us to find an orthogonal basis of a subspace. Let u 1,..., u k be a basis of a subspace V of R n. We begin the process of finding

More information

Lecture L3 - Vectors, Matrices and Coordinate Transformations

Lecture L3 - Vectors, Matrices and Coordinate Transformations S. Widnall 16.07 Dynamics Fall 2009 Lecture notes based on J. Peraire Version 2.0 Lecture L3 - Vectors, Matrices and Coordinate Transformations By using vectors and defining appropriate operations between

More information

Notes on Orthogonal and Symmetric Matrices MENU, Winter 2013

Notes on Orthogonal and Symmetric Matrices MENU, Winter 2013 Notes on Orthogonal and Symmetric Matrices MENU, Winter 201 These notes summarize the main properties and uses of orthogonal and symmetric matrices. We covered quite a bit of material regarding these topics,

More information

15.062 Data Mining: Algorithms and Applications Matrix Math Review

15.062 Data Mining: Algorithms and Applications Matrix Math Review .6 Data Mining: Algorithms and Applications Matrix Math Review The purpose of this document is to give a brief review of selected linear algebra concepts that will be useful for the course and to develop

More information

Au = = = 3u. Aw = = = 2w. so the action of A on u and w is very easy to picture: it simply amounts to a stretching by 3 and 2, respectively.

Au = = = 3u. Aw = = = 2w. so the action of A on u and w is very easy to picture: it simply amounts to a stretching by 3 and 2, respectively. Chapter 7 Eigenvalues and Eigenvectors In this last chapter of our exploration of Linear Algebra we will revisit eigenvalues and eigenvectors of matrices, concepts that were already introduced in Geometry

More information

CHAPTER 8 FACTOR EXTRACTION BY MATRIX FACTORING TECHNIQUES. From Exploratory Factor Analysis Ledyard R Tucker and Robert C.

CHAPTER 8 FACTOR EXTRACTION BY MATRIX FACTORING TECHNIQUES. From Exploratory Factor Analysis Ledyard R Tucker and Robert C. CHAPTER 8 FACTOR EXTRACTION BY MATRIX FACTORING TECHNIQUES From Exploratory Factor Analysis Ledyard R Tucker and Robert C MacCallum 1997 180 CHAPTER 8 FACTOR EXTRACTION BY MATRIX FACTORING TECHNIQUES In

More information

Solution to Homework 2

Solution to Homework 2 Solution to Homework 2 Olena Bormashenko September 23, 2011 Section 1.4: 1(a)(b)(i)(k), 4, 5, 14; Section 1.5: 1(a)(b)(c)(d)(e)(n), 2(a)(c), 13, 16, 17, 18, 27 Section 1.4 1. Compute the following, if

More information

1 Lecture 3: Operators in Quantum Mechanics

1 Lecture 3: Operators in Quantum Mechanics 1 Lecture 3: Operators in Quantum Mechanics 1.1 Basic notions of operator algebra. In the previous lectures we have met operators: ˆx and ˆp = i h they are called fundamental operators. Many operators

More information

2. Spin Chemistry and the Vector Model

2. Spin Chemistry and the Vector Model 2. Spin Chemistry and the Vector Model The story of magnetic resonance spectroscopy and intersystem crossing is essentially a choreography of the twisting motion which causes reorientation or rephasing

More information

Linear Algebra: Vectors

Linear Algebra: Vectors A Linear Algebra: Vectors A Appendix A: LINEAR ALGEBRA: VECTORS TABLE OF CONTENTS Page A Motivation A 3 A2 Vectors A 3 A2 Notational Conventions A 4 A22 Visualization A 5 A23 Special Vectors A 5 A3 Vector

More information

Linear Algebra I. Ronald van Luijk, 2012

Linear Algebra I. Ronald van Luijk, 2012 Linear Algebra I Ronald van Luijk, 2012 With many parts from Linear Algebra I by Michael Stoll, 2007 Contents 1. Vector spaces 3 1.1. Examples 3 1.2. Fields 4 1.3. The field of complex numbers. 6 1.4.

More information

The Quantum Harmonic Oscillator Stephen Webb

The Quantum Harmonic Oscillator Stephen Webb The Quantum Harmonic Oscillator Stephen Webb The Importance of the Harmonic Oscillator The quantum harmonic oscillator holds a unique importance in quantum mechanics, as it is both one of the few problems

More information

Inner Product Spaces

Inner Product Spaces Math 571 Inner Product Spaces 1. Preliminaries An inner product space is a vector space V along with a function, called an inner product which associates each pair of vectors u, v with a scalar u, v, and

More information

Mathematics Course 111: Algebra I Part IV: Vector Spaces

Mathematics Course 111: Algebra I Part IV: Vector Spaces Mathematics Course 111: Algebra I Part IV: Vector Spaces D. R. Wilkins Academic Year 1996-7 9 Vector Spaces A vector space over some field K is an algebraic structure consisting of a set V on which are

More information

Chapter 3. Cartesian Products and Relations. 3.1 Cartesian Products

Chapter 3. Cartesian Products and Relations. 3.1 Cartesian Products Chapter 3 Cartesian Products and Relations The material in this chapter is the first real encounter with abstraction. Relations are very general thing they are a special type of subset. After introducing

More information

The Characteristic Polynomial

The Characteristic Polynomial Physics 116A Winter 2011 The Characteristic Polynomial 1 Coefficients of the characteristic polynomial Consider the eigenvalue problem for an n n matrix A, A v = λ v, v 0 (1) The solution to this problem

More information

Inner Product Spaces and Orthogonality

Inner Product Spaces and Orthogonality Inner Product Spaces and Orthogonality week 3-4 Fall 2006 Dot product of R n The inner product or dot product of R n is a function, defined by u, v a b + a 2 b 2 + + a n b n for u a, a 2,, a n T, v b,

More information

PYTHAGOREAN TRIPLES KEITH CONRAD

PYTHAGOREAN TRIPLES KEITH CONRAD PYTHAGOREAN TRIPLES KEITH CONRAD 1. Introduction A Pythagorean triple is a triple of positive integers (a, b, c) where a + b = c. Examples include (3, 4, 5), (5, 1, 13), and (8, 15, 17). Below is an ancient

More information

a 11 x 1 + a 12 x 2 + + a 1n x n = b 1 a 21 x 1 + a 22 x 2 + + a 2n x n = b 2.

a 11 x 1 + a 12 x 2 + + a 1n x n = b 1 a 21 x 1 + a 22 x 2 + + a 2n x n = b 2. Chapter 1 LINEAR EQUATIONS 1.1 Introduction to linear equations A linear equation in n unknowns x 1, x,, x n is an equation of the form a 1 x 1 + a x + + a n x n = b, where a 1, a,..., a n, b are given

More information

Elements of Abstract Group Theory

Elements of Abstract Group Theory Chapter 2 Elements of Abstract Group Theory Mathematics is a game played according to certain simple rules with meaningless marks on paper. David Hilbert The importance of symmetry in physics, and for

More information

Vector and Matrix Norms

Vector and Matrix Norms Chapter 1 Vector and Matrix Norms 11 Vector Spaces Let F be a field (such as the real numbers, R, or complex numbers, C) with elements called scalars A Vector Space, V, over the field F is a non-empty

More information

by the matrix A results in a vector which is a reflection of the given

by the matrix A results in a vector which is a reflection of the given Eigenvalues & Eigenvectors Example Suppose Then So, geometrically, multiplying a vector in by the matrix A results in a vector which is a reflection of the given vector about the y-axis We observe that

More information

This unit will lay the groundwork for later units where the students will extend this knowledge to quadratic and exponential functions.

This unit will lay the groundwork for later units where the students will extend this knowledge to quadratic and exponential functions. Algebra I Overview View unit yearlong overview here Many of the concepts presented in Algebra I are progressions of concepts that were introduced in grades 6 through 8. The content presented in this course

More information

Figure 1.1 Vector A and Vector F

Figure 1.1 Vector A and Vector F CHAPTER I VECTOR QUANTITIES Quantities are anything which can be measured, and stated with number. Quantities in physics are divided into two types; scalar and vector quantities. Scalar quantities have

More information

13 MATH FACTS 101. 2 a = 1. 7. The elements of a vector have a graphical interpretation, which is particularly easy to see in two or three dimensions.

13 MATH FACTS 101. 2 a = 1. 7. The elements of a vector have a graphical interpretation, which is particularly easy to see in two or three dimensions. 3 MATH FACTS 0 3 MATH FACTS 3. Vectors 3.. Definition We use the overhead arrow to denote a column vector, i.e., a linear segment with a direction. For example, in three-space, we write a vector in terms

More information

Solving Systems of Linear Equations

Solving Systems of Linear Equations LECTURE 5 Solving Systems of Linear Equations Recall that we introduced the notion of matrices as a way of standardizing the expression of systems of linear equations In today s lecture I shall show how

More information

3. Mathematical Induction

3. Mathematical Induction 3. MATHEMATICAL INDUCTION 83 3. Mathematical Induction 3.1. First Principle of Mathematical Induction. Let P (n) be a predicate with domain of discourse (over) the natural numbers N = {0, 1,,...}. If (1)

More information

1 Sets and Set Notation.

1 Sets and Set Notation. LINEAR ALGEBRA MATH 27.6 SPRING 23 (COHEN) LECTURE NOTES Sets and Set Notation. Definition (Naive Definition of a Set). A set is any collection of objects, called the elements of that set. We will most

More information

NOTES ON LINEAR TRANSFORMATIONS

NOTES ON LINEAR TRANSFORMATIONS NOTES ON LINEAR TRANSFORMATIONS Definition 1. Let V and W be vector spaces. A function T : V W is a linear transformation from V to W if the following two properties hold. i T v + v = T v + T v for all

More information

Unified Lecture # 4 Vectors

Unified Lecture # 4 Vectors Fall 2005 Unified Lecture # 4 Vectors These notes were written by J. Peraire as a review of vectors for Dynamics 16.07. They have been adapted for Unified Engineering by R. Radovitzky. References [1] Feynmann,

More information

Metric Spaces. Chapter 7. 7.1. Metrics

Metric Spaces. Chapter 7. 7.1. Metrics Chapter 7 Metric Spaces A metric space is a set X that has a notion of the distance d(x, y) between every pair of points x, y X. The purpose of this chapter is to introduce metric spaces and give some

More information

Nonlinear Iterative Partial Least Squares Method

Nonlinear Iterative Partial Least Squares Method Numerical Methods for Determining Principal Component Analysis Abstract Factors Béchu, S., Richard-Plouet, M., Fernandez, V., Walton, J., and Fairley, N. (2016) Developments in numerical treatments for

More information

LS.6 Solution Matrices

LS.6 Solution Matrices LS.6 Solution Matrices In the literature, solutions to linear systems often are expressed using square matrices rather than vectors. You need to get used to the terminology. As before, we state the definitions

More information

Numerical Analysis Lecture Notes

Numerical Analysis Lecture Notes Numerical Analysis Lecture Notes Peter J. Olver 6. Eigenvalues and Singular Values In this section, we collect together the basic facts about eigenvalues and eigenvectors. From a geometrical viewpoint,

More information

Chapter 6. Orthogonality

Chapter 6. Orthogonality 6.3 Orthogonal Matrices 1 Chapter 6. Orthogonality 6.3 Orthogonal Matrices Definition 6.4. An n n matrix A is orthogonal if A T A = I. Note. We will see that the columns of an orthogonal matrix must be

More information

6. Vectors. 1 2009-2016 Scott Surgent (surgent@asu.edu)

6. Vectors. 1 2009-2016 Scott Surgent (surgent@asu.edu) 6. Vectors For purposes of applications in calculus and physics, a vector has both a direction and a magnitude (length), and is usually represented as an arrow. The start of the arrow is the vector s foot,

More information

Elasticity Theory Basics

Elasticity Theory Basics G22.3033-002: Topics in Computer Graphics: Lecture #7 Geometric Modeling New York University Elasticity Theory Basics Lecture #7: 20 October 2003 Lecturer: Denis Zorin Scribe: Adrian Secord, Yotam Gingold

More information

Linear Codes. Chapter 3. 3.1 Basics

Linear Codes. Chapter 3. 3.1 Basics Chapter 3 Linear Codes In order to define codes that we can encode and decode efficiently, we add more structure to the codespace. We shall be mainly interested in linear codes. A linear code of length

More information

Overview of Violations of the Basic Assumptions in the Classical Normal Linear Regression Model

Overview of Violations of the Basic Assumptions in the Classical Normal Linear Regression Model Overview of Violations of the Basic Assumptions in the Classical Normal Linear Regression Model 1 September 004 A. Introduction and assumptions The classical normal linear regression model can be written

More information

Applied Linear Algebra I Review page 1

Applied Linear Algebra I Review page 1 Applied Linear Algebra Review 1 I. Determinants A. Definition of a determinant 1. Using sum a. Permutations i. Sign of a permutation ii. Cycle 2. Uniqueness of the determinant function in terms of properties

More information

Chapter 17. Orthogonal Matrices and Symmetries of Space

Chapter 17. Orthogonal Matrices and Symmetries of Space Chapter 17. Orthogonal Matrices and Symmetries of Space Take a random matrix, say 1 3 A = 4 5 6, 7 8 9 and compare the lengths of e 1 and Ae 1. The vector e 1 has length 1, while Ae 1 = (1, 4, 7) has length

More information

26 Integers: Multiplication, Division, and Order

26 Integers: Multiplication, Division, and Order 26 Integers: Multiplication, Division, and Order Integer multiplication and division are extensions of whole number multiplication and division. In multiplying and dividing integers, the one new issue

More information

December 4, 2013 MATH 171 BASIC LINEAR ALGEBRA B. KITCHENS

December 4, 2013 MATH 171 BASIC LINEAR ALGEBRA B. KITCHENS December 4, 2013 MATH 171 BASIC LINEAR ALGEBRA B KITCHENS The equation 1 Lines in two-dimensional space (1) 2x y = 3 describes a line in two-dimensional space The coefficients of x and y in the equation

More information

1 VECTOR SPACES AND SUBSPACES

1 VECTOR SPACES AND SUBSPACES 1 VECTOR SPACES AND SUBSPACES What is a vector? Many are familiar with the concept of a vector as: Something which has magnitude and direction. an ordered pair or triple. a description for quantities such

More information

Partial Fractions. Combining fractions over a common denominator is a familiar operation from algebra:

Partial Fractions. Combining fractions over a common denominator is a familiar operation from algebra: Partial Fractions Combining fractions over a common denominator is a familiar operation from algebra: From the standpoint of integration, the left side of Equation 1 would be much easier to work with than

More information

Chapter 3. Distribution Problems. 3.1 The idea of a distribution. 3.1.1 The twenty-fold way

Chapter 3. Distribution Problems. 3.1 The idea of a distribution. 3.1.1 The twenty-fold way Chapter 3 Distribution Problems 3.1 The idea of a distribution Many of the problems we solved in Chapter 1 may be thought of as problems of distributing objects (such as pieces of fruit or ping-pong balls)

More information

Notes on Symmetric Matrices

Notes on Symmetric Matrices CPSC 536N: Randomized Algorithms 2011-12 Term 2 Notes on Symmetric Matrices Prof. Nick Harvey University of British Columbia 1 Symmetric Matrices We review some basic results concerning symmetric matrices.

More information

THE DIMENSION OF A VECTOR SPACE

THE DIMENSION OF A VECTOR SPACE THE DIMENSION OF A VECTOR SPACE KEITH CONRAD This handout is a supplementary discussion leading up to the definition of dimension and some of its basic properties. Let V be a vector space over a field

More information

Geometric Transformations

Geometric Transformations Geometric Transformations Definitions Def: f is a mapping (function) of a set A into a set B if for every element a of A there exists a unique element b of B that is paired with a; this pairing is denoted

More information

Vectors and Index Notation

Vectors and Index Notation Vectors and Index Notation Stephen R. Addison January 12, 2004 1 Basic Vector Review 1.1 Unit Vectors We will denote a unit vector with a superscript caret, thus â denotes a unit vector. â â = 1 If x is

More information

Structure of the Root Spaces for Simple Lie Algebras

Structure of the Root Spaces for Simple Lie Algebras Structure of the Root Spaces for Simple Lie Algebras I. Introduction A Cartan subalgebra, H, of a Lie algebra, G, is a subalgebra, H G, such that a. H is nilpotent, i.e., there is some n such that (H)

More information

Linear Algebra Review. Vectors

Linear Algebra Review. Vectors Linear Algebra Review By Tim K. Marks UCSD Borrows heavily from: Jana Kosecka kosecka@cs.gmu.edu http://cs.gmu.edu/~kosecka/cs682.html Virginia de Sa Cogsci 8F Linear Algebra review UCSD Vectors The length

More information

Section 6.1 - Inner Products and Norms

Section 6.1 - Inner Products and Norms Section 6.1 - Inner Products and Norms Definition. Let V be a vector space over F {R, C}. An inner product on V is a function that assigns, to every ordered pair of vectors x and y in V, a scalar in F,

More information

7 Gaussian Elimination and LU Factorization

7 Gaussian Elimination and LU Factorization 7 Gaussian Elimination and LU Factorization In this final section on matrix factorization methods for solving Ax = b we want to take a closer look at Gaussian elimination (probably the best known method

More information

In mathematics, there are four attainment targets: using and applying mathematics; number and algebra; shape, space and measures, and handling data.

In mathematics, there are four attainment targets: using and applying mathematics; number and algebra; shape, space and measures, and handling data. MATHEMATICS: THE LEVEL DESCRIPTIONS In mathematics, there are four attainment targets: using and applying mathematics; number and algebra; shape, space and measures, and handling data. Attainment target

More information

ASEN 3112 - Structures. MDOF Dynamic Systems. ASEN 3112 Lecture 1 Slide 1

ASEN 3112 - Structures. MDOF Dynamic Systems. ASEN 3112 Lecture 1 Slide 1 19 MDOF Dynamic Systems ASEN 3112 Lecture 1 Slide 1 A Two-DOF Mass-Spring-Dashpot Dynamic System Consider the lumped-parameter, mass-spring-dashpot dynamic system shown in the Figure. It has two point

More information

DRAFT. Further mathematics. GCE AS and A level subject content

DRAFT. Further mathematics. GCE AS and A level subject content Further mathematics GCE AS and A level subject content July 2014 s Introduction Purpose Aims and objectives Subject content Structure Background knowledge Overarching themes Use of technology Detailed

More information

Continuous Groups, Lie Groups, and Lie Algebras

Continuous Groups, Lie Groups, and Lie Algebras Chapter 7 Continuous Groups, Lie Groups, and Lie Algebras Zeno was concerned with three problems... These are the problem of the infinitesimal, the infinite, and continuity... Bertrand Russell The groups

More information

LINEAR INEQUALITIES. Mathematics is the art of saying many things in many different ways. MAXWELL

LINEAR INEQUALITIES. Mathematics is the art of saying many things in many different ways. MAXWELL Chapter 6 LINEAR INEQUALITIES 6.1 Introduction Mathematics is the art of saying many things in many different ways. MAXWELL In earlier classes, we have studied equations in one variable and two variables

More information

A Second Course in Mathematics Concepts for Elementary Teachers: Theory, Problems, and Solutions

A Second Course in Mathematics Concepts for Elementary Teachers: Theory, Problems, and Solutions A Second Course in Mathematics Concepts for Elementary Teachers: Theory, Problems, and Solutions Marcel B. Finan Arkansas Tech University c All Rights Reserved First Draft February 8, 2006 1 Contents 25

More information

Some Polynomial Theorems. John Kennedy Mathematics Department Santa Monica College 1900 Pico Blvd. Santa Monica, CA 90405 rkennedy@ix.netcom.

Some Polynomial Theorems. John Kennedy Mathematics Department Santa Monica College 1900 Pico Blvd. Santa Monica, CA 90405 rkennedy@ix.netcom. Some Polynomial Theorems by John Kennedy Mathematics Department Santa Monica College 1900 Pico Blvd. Santa Monica, CA 90405 rkennedy@ix.netcom.com This paper contains a collection of 31 theorems, lemmas,

More information

THE SIGN OF A PERMUTATION

THE SIGN OF A PERMUTATION THE SIGN OF A PERMUTATION KEITH CONRAD 1. Introduction Throughout this discussion, n 2. Any cycle in S n is a product of transpositions: the identity (1) is (12)(12), and a k-cycle with k 2 can be written

More information

PHY4604 Introduction to Quantum Mechanics Fall 2004 Practice Test 3 November 22, 2004

PHY4604 Introduction to Quantum Mechanics Fall 2004 Practice Test 3 November 22, 2004 PHY464 Introduction to Quantum Mechanics Fall 4 Practice Test 3 November, 4 These problems are similar but not identical to the actual test. One or two parts will actually show up.. Short answer. (a) Recall

More information

x1 x 2 x 3 y 1 y 2 y 3 x 1 y 2 x 2 y 1 0.

x1 x 2 x 3 y 1 y 2 y 3 x 1 y 2 x 2 y 1 0. Cross product 1 Chapter 7 Cross product We are getting ready to study integration in several variables. Until now we have been doing only differential calculus. One outcome of this study will be our ability

More information

Linearly Independent Sets and Linearly Dependent Sets

Linearly Independent Sets and Linearly Dependent Sets These notes closely follow the presentation of the material given in David C. Lay s textbook Linear Algebra and its Applications (3rd edition). These notes are intended primarily for in-class presentation

More information

Section 1.1. Introduction to R n

Section 1.1. Introduction to R n The Calculus of Functions of Several Variables Section. Introduction to R n Calculus is the study of functional relationships and how related quantities change with each other. In your first exposure to

More information

University of Lille I PC first year list of exercises n 7. Review

University of Lille I PC first year list of exercises n 7. Review University of Lille I PC first year list of exercises n 7 Review Exercise Solve the following systems in 4 different ways (by substitution, by the Gauss method, by inverting the matrix of coefficients

More information

Linear Programming. March 14, 2014

Linear Programming. March 14, 2014 Linear Programming March 1, 01 Parts of this introduction to linear programming were adapted from Chapter 9 of Introduction to Algorithms, Second Edition, by Cormen, Leiserson, Rivest and Stein [1]. 1

More information

LEARNING OBJECTIVES FOR THIS CHAPTER

LEARNING OBJECTIVES FOR THIS CHAPTER CHAPTER 2 American mathematician Paul Halmos (1916 2006), who in 1942 published the first modern linear algebra book. The title of Halmos s book was the same as the title of this chapter. Finite-Dimensional

More information

v w is orthogonal to both v and w. the three vectors v, w and v w form a right-handed set of vectors.

v w is orthogonal to both v and w. the three vectors v, w and v w form a right-handed set of vectors. 3. Cross product Definition 3.1. Let v and w be two vectors in R 3. The cross product of v and w, denoted v w, is the vector defined as follows: the length of v w is the area of the parallelogram with

More information

The Determinant: a Means to Calculate Volume

The Determinant: a Means to Calculate Volume The Determinant: a Means to Calculate Volume Bo Peng August 20, 2007 Abstract This paper gives a definition of the determinant and lists many of its well-known properties Volumes of parallelepipeds are

More information

Formal Languages and Automata Theory - Regular Expressions and Finite Automata -

Formal Languages and Automata Theory - Regular Expressions and Finite Automata - Formal Languages and Automata Theory - Regular Expressions and Finite Automata - Samarjit Chakraborty Computer Engineering and Networks Laboratory Swiss Federal Institute of Technology (ETH) Zürich March

More information

Cryptography and Network Security. Prof. D. Mukhopadhyay. Department of Computer Science and Engineering. Indian Institute of Technology, Kharagpur

Cryptography and Network Security. Prof. D. Mukhopadhyay. Department of Computer Science and Engineering. Indian Institute of Technology, Kharagpur Cryptography and Network Security Prof. D. Mukhopadhyay Department of Computer Science and Engineering Indian Institute of Technology, Kharagpur Module No. # 01 Lecture No. # 12 Block Cipher Standards

More information

CBE 6333, R. Levicky 1 Differential Balance Equations

CBE 6333, R. Levicky 1 Differential Balance Equations CBE 6333, R. Levicky 1 Differential Balance Equations We have previously derived integral balances for mass, momentum, and energy for a control volume. The control volume was assumed to be some large object,

More information

1 Symmetries of regular polyhedra

1 Symmetries of regular polyhedra 1230, notes 5 1 Symmetries of regular polyhedra Symmetry groups Recall: Group axioms: Suppose that (G, ) is a group and a, b, c are elements of G. Then (i) a b G (ii) (a b) c = a (b c) (iii) There is an

More information

Chapter 10 Electronic Wavefunctions Must Also Possess Proper Symmetry. These Include Angular Momentum and Point Group Symmetries

Chapter 10 Electronic Wavefunctions Must Also Possess Proper Symmetry. These Include Angular Momentum and Point Group Symmetries Chapter 10 Electronic Wavefunctions Must Also Possess Proper Symmetry. These Include Angular Momentum and Point Group Symmetries I. Angular Momentum Symmetry and Strategies for Angular Momentum Coupling

More information

Math 4310 Handout - Quotient Vector Spaces

Math 4310 Handout - Quotient Vector Spaces Math 4310 Handout - Quotient Vector Spaces Dan Collins The textbook defines a subspace of a vector space in Chapter 4, but it avoids ever discussing the notion of a quotient space. This is understandable

More information

Physics 235 Chapter 1. Chapter 1 Matrices, Vectors, and Vector Calculus

Physics 235 Chapter 1. Chapter 1 Matrices, Vectors, and Vector Calculus Chapter 1 Matrices, Vectors, and Vector Calculus In this chapter, we will focus on the mathematical tools required for the course. The main concepts that will be covered are: Coordinate transformations

More information

A note on companion matrices

A note on companion matrices Linear Algebra and its Applications 372 (2003) 325 33 www.elsevier.com/locate/laa A note on companion matrices Miroslav Fiedler Academy of Sciences of the Czech Republic Institute of Computer Science Pod

More information

How To Prove The Dirichlet Unit Theorem

How To Prove The Dirichlet Unit Theorem Chapter 6 The Dirichlet Unit Theorem As usual, we will be working in the ring B of algebraic integers of a number field L. Two factorizations of an element of B are regarded as essentially the same if

More information

5.61 Physical Chemistry 25 Helium Atom page 1 HELIUM ATOM

5.61 Physical Chemistry 25 Helium Atom page 1 HELIUM ATOM 5.6 Physical Chemistry 5 Helium Atom page HELIUM ATOM Now that we have treated the Hydrogen like atoms in some detail, we now proceed to discuss the next simplest system: the Helium atom. In this situation,

More information

2x + y = 3. Since the second equation is precisely the same as the first equation, it is enough to find x and y satisfying the system

2x + y = 3. Since the second equation is precisely the same as the first equation, it is enough to find x and y satisfying the system 1. Systems of linear equations We are interested in the solutions to systems of linear equations. A linear equation is of the form 3x 5y + 2z + w = 3. The key thing is that we don t multiply the variables

More information

5 Homogeneous systems

5 Homogeneous systems 5 Homogeneous systems Definition: A homogeneous (ho-mo-jeen -i-us) system of linear algebraic equations is one in which all the numbers on the right hand side are equal to : a x +... + a n x n =.. a m

More information

Arkansas Tech University MATH 4033: Elementary Modern Algebra Dr. Marcel B. Finan

Arkansas Tech University MATH 4033: Elementary Modern Algebra Dr. Marcel B. Finan Arkansas Tech University MATH 4033: Elementary Modern Algebra Dr. Marcel B. Finan 3 Binary Operations We are used to addition and multiplication of real numbers. These operations combine two real numbers

More information

Solution of Linear Systems

Solution of Linear Systems Chapter 3 Solution of Linear Systems In this chapter we study algorithms for possibly the most commonly occurring problem in scientific computing, the solution of linear systems of equations. We start

More information

Integrating algebraic fractions

Integrating algebraic fractions Integrating algebraic fractions Sometimes the integral of an algebraic fraction can be found by first epressing the algebraic fraction as the sum of its partial fractions. In this unit we will illustrate

More information